You are on page 1of 44

Accepted Manuscript

Heterogeneous photocatalytic removal of U(VI) in the presence of formic acid:


U(III) formation

Vanesa N. Salomone, Jorge M. Meichtry, Marta I. Litter

PII: S1385-8947(15)00155-2
DOI: http://dx.doi.org/10.1016/j.cej.2015.01.118
Reference: CEJ 13235

To appear in: Chemical Engineering Journal

Received Date: 13 November 2014


Revised Date: 26 January 2015
Accepted Date: 29 January 2015

Please cite this article as: V.N. Salomone, J.M. Meichtry, M.I. Litter, Heterogeneous photocatalytic removal of
U(VI) in the presence of formic acid: U(III) formation, Chemical Engineering Journal (2015), doi: http://dx.doi.org/
10.1016/j.cej.2015.01.118

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Heterogeneous photocatalytic removal of U(VI) in the presence of

formic acid: U(III) formation

Vanesa N. Salomonea, Jorge M. Meichtrya,b, Marta I. Littera,b,c*

a
Gerencia Qumica, Comisin Nacional de Energa Atmica, Av. Gral. Paz 1499, 1650

San Martn, Prov. de Buenos Aires, Argentina.


b
Consejo Nacional de Investigaciones Cientficas y Tcnicas (CONICET), Av.

Rivadavia 1917, 1033 Ciudad Autnoma de Buenos Aires, Argentina


c
Instituto de Investigacin e Ingeniera Ambiental, Universidad Nacional de General

San Martn, Campus Miguelete, Av. 25 de Mayo y Francia, 1650 San Martn, Prov. de

Buenos Aires, Argentina

Abstract

The efficiency of TiO2 heterogeneous photocatalysis under UV-visible light for

removal of uranyl (UO22+) (0.25 mM, pH 3) in the presence of formic acid (HCOOH)

was investigated. The effect of the counterion of the uranium salt (perchlorate and

acetate-nitrate) and the use of quartz and glass photoreactors were analyzed. The

transformation of U(VI) in the presence of HCOOH was significantly high and the best

removal conditions have been established. The highest efficiency was observed for

*
Corresponding author. Tel.: +541167727016; fax: +541167727886.E-mail addresses:
litter@cnea.gov.ar; marta.litter@gmail.com (M.I. Litter).

1
uranyl perchlorate in the quartz photoreactor with 0.001 M HCOOH, while uranyl

acetate yielded a lower removal. HCOOH concentrations > 0.01 M provoked uranium

reoxidation. U(VI) photocatalytic transformation was also studied by direct

spectrophotometry using TiO2 nanoparticles in the presence of 1 M HCOOH at pH < 2.

U(V), U(IV) and U(III) were detected. U(III) formation was also apparent in the

absence of the photocatalyst. This is the first time that formation of U(III) in a

photochemical U(VI) system is reported.

Keywords

Heterogeneous Photocatalysis, U(VI), U(III), Formic Acid

1. Introduction

Heterogeneous photocatalysis (HP) of aqueous U(VI) systems has been rather well

studied [1-5] as an economical and simple alternative to conventional removal methods

[6-8]. This technological approach can solve not only environmental problems but also

those related to actinide valence state control for waste minimization in nuclear fuel

processing. However, studies on the effect of several variables such as the nature of

electron donors or the presence of some ions in the system should be completed because

they have profound effects on the efficiency of the transformation. In a similar previous

paper [9], the uranyl HP system has been studied in the presence of 2-propanol (PrOH).

2
High U(VI) removal efficiencies were obtained, with complete removal for uranyl

nitrate and uranyl perchlorate with 1 M 2-PrOH in a quartz photoreactor. The acetate

salt showed a lower U(VI) removal. It was proposed that U(VI) reduction is mediated

by conduction band electrons (eCB) and not by the organic radicals formed by hole/HO

attack to 2-PrOH [9]. The homogeneous photochemical uranyl system (absence of TiO2)

has also been revisited in a previous work [10]. Although the photochemical reaction

under the same conditions to those of the photocatalytic one attained a similar

efficiency, the HP treatment exhibited a better performance.

The uranyl HP transformation in the presence of different organic compounds,

formic acid (HCOOH) among them, has been studied by Amadelli et al. in an early

work with TiO2 P25, and it was found that the photoelectrochemical conversion rate

followed the order formate > acetate > 2-PrOH [1].

The present work aimed to revisit the uranyl HP system in the presence of HCOOH

with P25 at pH 3, analyzing the effects of the counterion of the uranyl salt (perchlorate

and acetate) and the use of quartz and glass photoreactors. Taking into account that in

the presence of 2-PrOH [9] uranyl perchlorate and nitrate reacted similarly, the nitrate

salt was not tested in the present work. Experiments with transparent TiO2 nanoparticles

(pH < 2) were also performed to shed light on the involved mechanisms.

3
2. Experimental

2.1. Materials and chemicals

Uranyl acetate (UO2(CH3COO)22H2O, Fluka), uranyl nitrate (UO2(NO3)26H2O,

Lopal) and HCOOH (99%, Carlo Erba) were used. Uranyl perchlorate was prepared

from uranyl nitrate and HClO4(c) (Merck), following a previous procedure [4]. Uranyl

acetate solutions were prepared by dissolving uranyl acetate in 1% HNO3. TiO2 (now

AEROXIDE TiO2 P25, Evonik) was provided by Degussa (Germany) and used as

received. All other chemicals were reagent grade and used without further purification.

Water was purified with a Millipore Milli-Q equipment (resistivity = 18 Mcm).

TiO2 nanoparticles were synthesized in the laboratory, according to a literature

procedure [11]. Briefly, 200 mL of water at pH 1.5 (HClO4) were placed in an ice bath

and a solution of 1 mL of titanium isopropoxide in 20 mL of 2-PrOH was slowly added

with constant stirring. The final mixture was stirred for 2 to 3 days in the ice bath until a

clear solution was obtained. TiO2 concentration (0.012 M, 1 g L1) was calculated from

the used quantity of titanium isopropoxide.

2.2. Irradiation experiments

Photocatalytic experiments with TiO2 suspensions were carried out with an

irradiation setup consisting of a commercial quartz (Q) photoreactor immersion well

4
(Photochemical Reactors Ltd.) provided with a medium pressure mercury lamp (125 W,

> 230 nm, max = 365 nm), surrounded by a thermostatic jacket, set at 25 C and

acting as IR filter (setup I). Other emissions of the lamp were at 245, 254, 265, 280,

302, 313, 408, 436 and 546 nm. The incident photon flux per unit volume (q0n,p/V),

measured by actinometry with potassium ferrioxalate, was 121 einstein s1 L1. In

selected experiments, a glass well was used (G photoreactor, > 310 nm, max = 365

nm, q 0n,p/V = 44 einstein s1 L1).

TiO2 suspensions (200 mL, 1 g L1) containing 0.25 mM uranyl were irradiated

under nitrogen bubbling (0.5 L min1). When indicated, the corresponding volume of

HCOOH was added to reach the corresponding concentration (0.00025 to 1 M). pH was

adjusted to 3 with 2 M NaOH or with concentrated HClO4 (70%). The following

conditions were used in the irradiation experiments: i) uranyl perchlorate with

perchloric acid, quartz photoreactor, hereafter named QP and ii) uranyl acetate in nitric

acid, quartz photoreactor, QAN. The experiments with uranyl perchlorate and the glass

photoreactor system will be named GP. In QAN, initial nitrate and acetate

concentrations were 22.0 and 0.5 mM, respectively; in QP and GP, initial perchlorate

concentration was 1.5 mM. Some experiments were carried out with the reactor open to

the air or under air bubbling (0.5 L min1). Before switching on the lamp, the

suspension was carefully stirred in the dark during 30 min to ensure the adsorption

equilibrium of U(VI) and HCOOH onto TiO2, and to evaluate concentration changes

only under irradiation. Aliquots (250 L) were periodically taken during the runs,

filtered through Millipore membranes (0.2 m) or centrifuged (see below) before

5
analysis. Deposits were carefully conserved under vacuum before analysis. In the first

experiments, the samples were filtered using only one membrane for all samples of the

run; in this case, leaching of U(VI) to the solution was observed in samples taken at 10

min and later, detected by an increase of the U(VI) concentration in solution due to

oxidation of the formed U(IV) under air contact. To avoid this leaching, in further runs,

one membrane per each sample was used or, alternatively, the sample was rapidly

centrifuged. Changes of pH in all runs were negligible (pH < 0.3). Reactions in the

absence of TiO2 were performed under identical conditions.

All the experiments were performed at least twice and the results were averaged. The

experimental error was never higher than 10%, as calculated by standard deviation

among the replicate experiments; error bars for the averaged experiments are shown in

the corresponding figures. The fitting of experimental points was performed with Origin

8.0 software.

The irradiation setup for the experiments with TiO2 nanoparticles consisted of a

cylindrical cell (10 mL, 1 cm pathlength) with quartz walls irradiated with a 150 W

ozone-free xenon lamp (Newport) (setup II). Continuous wavelengths from the UV-C

up to the IR were emitted from the lamp. The UV-A irradiance, measured with a

Spectroline DM-365 XA radiometer placed at 10 cm from the lamp, was E0 = 3700 W

cm2. These photocatalytic runs were performed with uranyl perchlorate dissolved in

0.88% perchloric acid, the corresponding amount of HCOOH, and 0.010 M TiO2

nanoparticles. N2 was bubbled for 10 min (0.1 L min1) before the irradiation and then

6
the cell was closed. Experiments in the absence of TiO2 were performed under the same

conditions.

2.3. Analytical determinations

U(VI) concentration was followed by the spectrophotometric PAR technique [12],

with a detection limit of 0.25 M uranyl. The solid deposits were analyzed by total

reflection X-ray fluorescence (TXRF, modular equipment, Seifert X-ray generator, with

a Canberra detector). For XRD analyses, a Philips PW-3710 diffractometer was used.

In the experiments with nanoparticles, the temporal evolution of uranium signals was

followed by direct UV spectroscopy using a Hewlett Packard 8453 spectrophotometer.

3. Results

3.1. Heterogeneous photocatalytic experiments

The temporal evolution of normalized U(VI) concentration during HP experiments in

anoxic conditions starting from uranyl perchlorate and two HCOOH concentrations

(0.001 and 1 M) up to 120 min irradiation is shown in Fig. 1. In the absence of

HCOOH, final uranyl removal values were 43 and 24% for the Q and G photoreactors,

respectively, and no precipitates were observed on the TiO2 surface. In the presence of

0.001 M HCOOH, removal was higher, reaching percentages above 90% without

7
significant differences between both photoreactors and with very similar kinetic decay

profiles; dark grey precipitates appeared on the photocatalyst surface. At 1 M HCOOH,

removal was lower for both photoreactors, especially for G; with this photoreactor, after

an initial rapid U(VI) decrease, a significant increase in the concentration of U(VI) in

solution took place after 5 min. Interestingly, no precipitates on TiO2 were formed, but a

brick red color was visible in solution at the end of the experiments under both

conditions, which could be attributed to U(III) formation, as we will describe later, and

which disappeared very fast when the sample was filtered in contact with air. Similar

experiments in the absence of the photocatalyst with 1 M HCOOH and both

photoreactors indicated a lower removal compared with the HP experiments, with a

higher removal with the Q photoreactor (63 and 14%, respectively for QP and GP),

similarly to the system with 2-PrOH [10]. However, in contrast with this system, where

dark and light yellow precipitates have been observed, no precipitate was formed,

suggesting that U(IV) species stay in solution as HCOOH complexes. No U(VI)

removal was observed in the dark in any case, showing that uranyl adsorption over TiO2

is negligible at the working pH, according to previously reported data [9].

8
1.0

0.8
[U(VI)]/[U(VI)]0

0.6
QP, no HCOOH
QP, [HCOOH] = 0.001 M
QP, [HCOOH] = 1 M
0.4 GP, no HCOOH
GP, [HCOOH] = 0.001 M
GP, [HCOOH] = 1 M
QP, no TiO2, [HCOOH] = 1 M
0.2 GP, no TiO2, [HCOOH] = 1 M

0.0
0 20 40 60 80 100 120
Time (min)

Fig. 1. Temporal evolution of normalized U(VI) concentration in the presence of

HCOOH during HP experiments with uranyl perchlorate. Conditions: setup I, QP and

GP, [U(VI)]0 = 0.25 mM, [TiO2] = 1 g L1, N2 (0.5 L min1), pH 3, T = 25 C. Q

photoreactor: > 230 nm, max = 365 nm, q 0n,p/V = 121 einstein s1 L1. G

photoreactor: > 310 nm, max = 365 nm, q0n,p/V = 44 einstein s1 L1. Dotted lines are

fittings of experimental points to Eq. (1), with the exception of GP with 1 M HCOOH,

which was fitted up to 5 min, and experiments without TiO2, where dotted lines are only

for a better visualization of the experimental points.

Fig. 2 shows results of similar experiments starting from uranyl acetate in nitric acid

in the Q photoreactor (QAN) and various HCOOH concentrations (0.00025 - 1 M).

Experiments were run up to 120 min but no changes were observed after 30 min. A

significant and rapid removal (60 and 80%) in the first 7 min took place for all

9
conditions with the exception of the lowest HCOOH concentration (0.00025 M), which

presented a slower initial decay. In the experiments without HCOOH and with HCOOH

up to 0.01 M, after this rapid decay, uranyl removal stopped without changes up to the

end of the run. In the runs at the highest HCOOH concentrations (0.1 and 1 M), an

increase in the U(VI) concentration took place after 7 min. Similarly as observed for QP

and GP at 1 M HCOOH, no precipitate was formed over TiO2, and a brick red color was

observed in the solution at the end of the experiment, which disappeared by contact with

O2. No uranyl decay was observed in a run with 0.1 M HCOOH in the absence of light.

A run under air bubbling (0.1 M HCOOH) yielded a U(VI) decay of only 10% at 30

min, followed by an arrest; a similar run with the reactor open to air gave a somewhat

higher decay, 50% in 30 min.

HP, no HCOOH
HP, [HCOOH] = 0.00025 M
HP, [HCOOH] = 0.001 M
HP, [HCOOH] = 0.01 M
HP, [HCOOH] = 0.1 M
HP, [HCOOH] = 1 M
1.0
HP, no light, [HCOOH] = 0.1 M
HP, no N 2 , [HCOOH] = 0.1 M
HP, with O2, [HCOOH] = 0.1 M
0.8
[U(VI)]/[U(VI)]0

0.6

0.4

0.2

0.0
0 5 10 15 20 25 30

Time (min)

10
Fig. 2. Temporal evolution of normalized U(VI) concentration in the presence of

HCOOH during experiments with TiO2 with uranyl acetate (QAN conditions).

Experiments were run up to 120 min but no changes were observed after 30 min.

Conditions of Fig. 1. Dotted lines are fittings of experimental points to Eq. (1) with the

exception of the curves with 0.1 and 1 M HCOOH, which were fitted up to 5 min.

Results of all kinetic profiles for the HP experiments could be adjusted to

monoexponential decays, according to Eq. (1):

[ U(VI) ]t
= A exp k t + (1 A) (1)
[ U(VI) ]0

where A is the fraction of U(VI) removed by the photocatalytic process and k is the

pseudo first order kinetic constant. In the experiments at the highest HCOOH

concentrations, where a redissolution of U(VI) was observed and, only for comparison,

an adjustment with Eq. (1) was made for the first five minutes.

Table 1 presents the comparative results of all experiments indicating the extent of

uranyl removal at 120 min (at 30 min for experiments for QAN), the type of precipitate

and the values of kinetic parameters with their errors (obtained from the fittings). The

fittings to the monoexponential regime were very good in all cases (R2 0.90) and

reflect the profiles of Figs. 1 and 2. The discussion of the values in the absence of

donors has been already reported [9], with A decreasing in the order QP > GP > QAN

and k indicating a very slow initial removal, with a higher value for QAN. In the

11
presence of HCOOH, both A and k increased under all conditions, except for

QAN/0.00025 M HCOOH, where A was similar to the experiment without HCOOH.

For the perchlorate system, the highest removal and A values were obtained at

[HCOOH] = 0.001 M in both photoreactors, but a higher k value was obtained at

[HCOOH] = 1 M. The use of the quartz photoreactor gave better results than the use of

the glass one, with the exception of 0.001 M HCOOH, where A and k were similar. For

the QAN system, the optimal HCOOH concentration was 0.01 M, where the higher

removal was reached. An increasing HCOOH concentration caused a decrease in U(VI)

removal, ascribed to the formation of soluble formate complexes [13], followed by

U(IV) reoxidation during filtration [9]. The values of A and k were almost constant for

all QAN systems at [HCOOH] 0.01 M. Except for GP at 1 M HCOOH, the percentage

of removal and A were equal or smaller for QAN, indicating a detrimental effect of

acetate, as previously observed with 2-PrOH [9]. A decrease of A and k was observed in

QAN experiments with the reactor open to air or under air bubbling. The kinetic

parameters point out that a smaller fraction of U(VI) was removed at a reduced rate, the

results being much more noticeable under air bubbling due to the higher O2

concentration.

12
Table 1

Percentage of uranyl removal at 120 min, type of precipitate obtained at the end of the run and kinetic parameters obtained from

the fitting of the experimental points of Figs. 1 and 2 with Eq. (1).

Condition/[HCOOH] (M) % U(VI) Precipitate A k (min1) R2

removal

TiO2, QP, 0 43 No 0.56 0.08 0.013 0.003 0.98

TiO2, GP/0 24 No 0.45 0.04 0.006 0.002 0.90

TiO2, QAN/0 26a No 0.32 0.03 0.068 0.005 0.95

TiO2, QP/0.001 94 Yes (dark grey) 0.90 0.03 0.19 0.02 0.99

TiO2, GP/0.001 90 Yes (dark grey) 0.87 0.05 0.17 0.01 0.98

TiO2, QP/1 78 No 0.78 0.09 0.37 0.08 0.91

TiO2, GP/1b 29 No 0.47 0.04 0.4 0.1 0.97

No TiO2, QP/1 63 No ND ND ND

No TiO2, GP/1 14 No ND ND ND

13
TiO2, QAN/0.00025 44a No 0.40 0.03 0.26 0.02 0.96

TiO2, QAN/0.001 75a Yes (light grey) 0.74 0.07 0.34 0.01 0.99

TiO2, QAN/0.01 79a Yes (light grey) 0.81 0.02 0.50 0.08 0.99

TiO2, QAN/0.1b 49a No 0.81 0.02 0.5 0.1 0.99

TiO2, QAN/1b 43a No 0.75 0.02 0.50 0.04 0.99

TiO2, no light, QAN/0.1 0a No ND ND ND

TiO2, open to air, QAN/0.1 51a No 0.53 0.02 0.10 0.01 0.99

TiO2, with O2, QAN/0.1 11a No 0.11 0.03 0.10 0.01 0.90

a
At 30 min
b
Adjustment only up to the first five min

ND: not determined

14
Remarkably, the formation of precipitates was observed only at 0.001 and 0.01 M

HCOOH, either with QP or with QAN. At lower HCOOH concentrations, not enough U

removal was probably achieved, and, as said, at higher HCOOH concentrations, soluble

complexes are formed [13]. Dark grey precipitates were observed for QP/0.001 and

GP/0.001, while for QAN/0.001 and QAN/0.01, precipitates were light grey, in

agreement with the lower U(VI) removal. The color suggested the formation of oxides

of stoichiometry UO2+x (x = 0-0.25) [1,9], with a predominance of U(IV). Although

TXRF analysis of these deposits (not shown) confirmed the presence of uranium, XRD

analysis reported only signals of TiO2, probably due to the low uranium concentration

or to the lack of crystallinity of deposits. Thus, these analyses did not allow the

identification of the uranium oxides.

A comparison of the absorbance spectra of the filtered suspensions before and after a

HP experiment for QAN with 1 M HCOOH is presented in Fig. 3. The initial spectrum

showed the characteristic U(VI) signals in the region 350-500 nm (see inset) [14-16],

while the final one showed U(IV) signals at 435, 490, 554 and 663 nm together with a

broad band at 800-900 nm [17-21]. This result confirms that uranium remains in

solution in the form of a U(IV)-formate complex [13], being reoxidized to U(VI) by O2

during the filtration and increasing U(VI) concentration in solution after approximately

7 min (Fig. 2). The peak at 963 nm indicates that U(V) is also formed [22,24].

15
0.05 4 0.008

3 0.006

Absorbance

Absorbance
0.04
2 0.004

Absorbance
1 0.002
0.03

0 0.000
200 250 300 350 350 375 400 425 450 475 500

0.02 Wavelength (nm) Wavelength (nm)

t = 0 min
t = 120 min

0.01

0.00
300 400 500 600 700 800 900 1000

Wavelength (nm)

Fig. 3. Spectra of the initial and final filtered solutions of a HP experiment for QAN in

the presence of 1 M HCOOH. Conditions of Fig. 1. Insets: detailed views of the 200-

350 nm and 350-500 nm regions.

3.2. Irradiation experiments at pH < 2 in the presence of TiO2 nanoparticles

Irradiation experiments of 1 mM uranyl perchlorate solutions with 0.010 M TiO2

nanoparticles in the absence and in the presence of 1 M HCOOH at very acid pH (< 2)

under N2 were performed using setup II. Higher initial U(VI) concentrations than in the

experiments with P25 suspensions were used to clearly observe changes. Fig. 4 shows

the evolution of the UV-Vis spectra of the uranyl perchlorate solution with the TiO2

nanoparticles and absence of HCOOH after 120 min irradiation. After this time, no

more spectral changes were observed.

16
0.03
0.05

0.04

Abs orbance
0.03

0.02 0.02
Absorbance

0.01

0.00
350 400 450 500
Wave le ngth (nm )
0.01

t = 0 min
t = 120 min
no U(VI)

0.00
300 400 500 600 700 800 900 1000

Wavelength (nm)

Fig. 4. Temporal evolution of the spectrum of a uranyl perchlorate solution in the

presence of TiO2 nanoparticles and absence of HCOOH. Conditions: setup II, [U(VI)]0

= 1 mM, [TiO2] = 0.010 M, pH 1.80, N2 atmosphere, E0 = 3700 W cm2, full emission

spectrum. The spectrum in the absence of U(VI) is also included. Inset: detailed view of

the 350-500 nm region.

The spectrum of the initial solution shows the characteristic U(VI) signals (see inset)

[14-16] overlapped with the spectrum of the TiO2 nanoparticles. The signals of U(VI)

disappeared after irradiation, indicating a complete reduction, and three new signals at

737, 836 and 963 nm appeared, corresponding to U(V) [22-24]. The spectrum of the

nanoparticles in the absence of U(VI) is included, showing only the TiO2 absorbance in

the 250-400 nm range [25].

17
Fig. 5 shows the initial and final spectra of a HP run under the same conditions of

Fig. 4 but in the presence of 1 M HCOOH. The initial spectrum shows signals of U(VI)

[15]. After 120 min, peaks of U(IV) at 663 and 800-900 nm [17-21], and of U(V) at 737

and 963 nm [22-24] can be seen; the peak at 836 nm visible in Fig. 4 was surely

overlapped by the broad U(IV) band. Remarkably, an important increase in the

absorbance in the 350-600 nm range was observed (see inset), together with a change of

color in the suspension, from transparent to deep orange, which coincides with the

reported absorbance of U(III) species [19,20,26,27], corresponding probably to a U(III)-

HCOOH complex [28-30]. Other peaks corresponding to U(III) were not observed,

surely due to a masking by the strong U(IV) signals [19,20].

0.05
0.5

0.4
Absorbance

0.04 0.3

0.2

0.1
Absorbance

0.03
0.0
350 400 450 500 550 600
Wave le ngth (nm )

0.02

t=0
0.01 t = 120 min
no U(VI)

0.00
300 400 500 600 700 800 900 1000

Wavelength (nm)

Fig. 5. Temporal evolution of the spectrum of a uranyl perchlorate solution in the

presence of TiO2 nanoparticles and HCOOH. Conditions: setup II, [U(VI)]0 = 1 mM,

18
[TiO2] = 0.010 M, [HCOOH] = 1 M, pH 1.3, N2 atmosphere, E0 = 3700 W cm2, full

emission spectrum. Inset: detailed view of the 350-600 nm region.

Results of similar experiments in the absence of TiO2 nanoparticles are shown in Fig.

6, where the spectra of the initial solution and after 90 min of irradiation can be seen.

The spectrum of the initial sample shows the characteristic U(VI) peaks at 350-500 nm

(see inset) [15]. After 90 min of irradiation, the spectrum showed signals of U(V) (963

nm) [22-24] and U(IV) (450, 500, 550, 650 and 850 nm) [17-21]; as no further changes

were observed and the spectrum was very different from that of U(VI), it will be

assumed that U(VI) was almost completely reduced. At 90 min, the cell was opened to

air and left in the dark for 30 min more. The spectrum registered the disappearance of

the 963 nm peak due certainly to the fast oxidation or disproportionation of U(V)

[24,31,32-35]. Interestingly, a small decrease of the absorbance in the range 350-600

nm can be observed, which cannot be ascribed neither to U(V) [22-24] nor to U(IV)

[17-21]. After the reaction with O2, if this absorbance would have corresponded to

U(VI), an increase in the absorbance should have been observed; therefore, the decrease

in the absorbance can only be ascribed to U(III) formation. In addition, the U(IV) peak

centered at 850 nm was certainly higher, while the peak at 650 nm seemed to be more

defined. A very similar change of the U(IV) spectra during U(VI) radiolytic studies was

found by Elliot et al., who also ascribed this change to U(III) [36]. Although U(III) is

unstable, its half-life has been reported to be at least 755 h under similar conditions to

ours [37], enough to be spectroscopically detected.

19
0.10
t=0 0.10
0.09 t = 90 min
t = 90 min + 30 min in air, the dark 0.08

Absor bance
0.08 0.06

Absorbance 0.07 0.04

0.02
0.06
0.00
0.05 350 400 450 500 550 600
Wavele ngth (nm )
0.04

0.03

0.02

0.01

0.00
300 400 500 600 700 800 900 1000
Wavelength (nm)

Fig. 6. Temporal evolution of the spectrum of a uranyl perchlorate solution in the

presence of HCOOH (no TiO2 nanoparticles) up to 90 min of UV-Vis irradiation and

after 90 min of UV-Vis irradiation plus 30 min of exposure in the dark to the air.

Conditions: setup II, [U(VI)]0 = 1 mM, [HCOOH] = 1 M, pH 1.9, N2 atmosphere, E0 =

3700 W cm2, full emission spectrum. Inset: detailed view of the 350-600 nm region.

4. Discussion

For P25, the values of the edges of conduction band (CB) and valence band (VB) at

pH 0 have been calculated as 0.3 and +2.9 V vs. SHE, respectively [38]. Under

normal laboratory UV illumination conditions and at pH 3, P25 eCB are able to

photocatalytically transform UO22+ into UO2+ and U(IV) through one-electron

All reduction potentials given in this paper are standard values vs. SHE.

20
consecutive steps (Eq. (2)), while further reduction to other uranium oxidation states is

not possible, according to the Latimer diagram (Scheme 1).

Scheme 1. Latimer diagram for different uranium species vs. SHE [27,39-41].

The discussion of results in the absence of electron donors can be reviewed in the

previous work [9]. Briefly, uranyl removal from water takes place, but it is low because

the conjugate reaction, oxidation of water by hVB+ to give free HO (Eq. (3)) is slow

(yielding finally H2O2 and O2), and reoxidation to U(V/VI) by hVB+ or HO (Eq. (4))

competes, leading to a short-circuit and fast e/hVB+ recombination, both stopping U(VI)

transformation.

U(VI)/U(V) + eCB U(V)/U(IV) (2)

H2O + hVB+ HO + H+ (3)

U(IV)/U(V) + hVB+/HO U(V)/U(VI) (4)

The reported characteristic time for the formation of trapped HO on TiO2 according

to Eq. (3) is 10 ns [42]. Formation of HO is a necessary and well-known reaction in

21
photocatalytic systems, which induces oxidation of different species present in the

medium [42].

Although no values were found for the reaction of U(IV) or U(V) with TiO2-hVB+,

the reported value for the reaction of U(IV) with free HO in aqueous solution ranges

between 8 108 M1 s1 [36] and 8.6 108 M1 s1 [43,44]. For the UO2+/HO radical

pair in water, a lifetime of 20 ns [31 and references therein] was found. This means that

both reduced U species are very rapidly oxidized by hydroxyl radicals.

The faster reaction at initial times for QAN in the absence of donor was explained

because acetate acts as a donor, favoring uranyl transformation; however, at longer

times, U(VI) concentration in solution increases due to the formation of soluble and

oxidizable U(IV)-acetate complexes, as reflected by the smaller value of A (Table 1)

[9].

In the presence of electron donors at high concentrations, recombination of eCB/hVB+

pairs and reoxidation of uranium species is prevented, since these species are oxidized

by hVB+ or HO in preference to uranium species. Moreover, the sacrificial donor

generates highly reactive radicals by attack of hVB+ or HO (E0 = +2.8 V for HO/H2O)

[45], which enhance reductive steps.

Regarding HCOOH addition as electron donor, it is interesting to analyze first the

effect of this electron donor in the absence of TiO2. Several authors examined the

photochemical reaction of uranyl in the presence of HCOOH [46-53], and some of them

[46,50] proposed the following process, with U(IV) and CO2 being the only detectable

products and U(V) an intermediate species.

22
UO22+ + 3 H+ + HCOO + h U4+ + CO2 + 2 H2O (5)

In a recent work, Lucks et al. [54] studied the U(VI) photoreduction in the presence

of HCOOH through density functional theory calculations, and concluded that the

reaction takes place by intermolecular hydrogen abstraction only from the protonated

acid, as also proposed by McCleskey et al. some years before [53]. According to Lucks

et al. [54], the abstracted H is the one linked to the C atom, with generation of

COOH/COO radicals (as pKa = 1.4, the COO anion is the dominant species at the

working pH) [55].

UO22+ + h [UO22+]* (6)

[UO22+]* + HCOO UO2 + + COO + H+ (7)

A second order rate constant around 2 107 M1 s1 was determined for Eq. (7),

while the lifetime of [UO22+]* in the absence of chemical quenchers was 50 s [53],

both in water.

Considering E0 for CO2 1.9 V [45], this radical is able to reduce U(VI) to U(V)

and U(IV) but also to U(III) (Eq. (8)). Even U(II) might be generated, but its existence

is still unclear [27,56] and it will not be considered here.

U(VI)/U(V)/U(IV) + COOH/COO U(V)/U(IV)/U(III) + CO2 (8)

23
The possible reoxidation of U(V) by COO/COOH (Eq. (9)), proposed elsewhere

[53], cannot be completely discarded, due to the acidic conditions [57]. However, this

would cause only consumption of HCOOH without net U(VI) reduction.

U(V) + COOH/COO U(VI) + CO (9)

U(III) can be oxidized by hVB+ or HO:

U(III) + hVB+/HO U(IV) (10)

The rate constant for the reaction of U(III) with hVB+ is unknown, but a value of 4.1

108 M1 s1 has been reported for the reaction with HO [58]. It can be suggested that

U(III) can be easily oxidized by hBV+/HO.

In the presence of TiO2, the process will be similar but mediated by the

semiconductor. First, direct U(VI) reduction by eCB (Eq. (2)) takes place up to U(IV).

The conjugate reaction is generation of COOH/COO by reaction (11), ending in CO2

through further oxidation (Eq. (12)) [59].

HCOOH/HCOO + hVB+/HO COOH/COO + H2O (11)



COOH/COO + hVB+/HO CO2 + H2O (12)

24
The rate constant of HCOOH with HO in water is reported to be 1.3 108 M1 s1

[43], while for the formate ion is 3.2 109 M1 s1 [43,59]. Similarly to other electron

donors, HCOOH 1) accelerates uranyl reduction by reducing the recombination rate of

eCB/hVB+ pairs, and 2) is able to indirectly reduce U(VI) by COO [36]. Photochemical

experiments at acid pH (< 2) with TiO2 nanoparticles in the presence of high HCOOH

concentrations (Fig. 5) allowed to elucidate and support the mechanism of HP uranyl

reduction by direct detection of species in solution. The anoxic and acidic pH conditions

used were appropriate for detection of U(III) [19,37], but this occurred significantly

only in the presence of TiO2 nanoparticles and HCOOH. In this case, all reduced

uranium species, i.e., U(V), U(IV) and U(III) could be detected. The change of color of

the suspension also confirmed U(III) formation. Interestingly, U(III) seems to have been

formed also in the HP experiments with P25 at high HCOOH concentrations (pH 3) in

QP, GP and QAN, as suggested by the brick red color developed in the suspension [27]

which disappeared in air [19,27,60]. This contrasts with the 2-PrOH system [9], where

this change of color was never appreciated, and neither U(IV) nor U(III) in solution was

observed. Although U(IV) can be produced by direct eCB reduction, it is not detected in

solution in the 2-PrOH system due to the fast deposition on the TiO2 surface as UO2+x.

In the previous 2-PrOH study, when starting from uranyl acetate [9], it was shown that

even when acetate was able to solubilize U(IV), U(III) was not formed either. In

HCOOH, U(IV) stays in solution as a formate complex [13], allowing further reduction

to U(III). On the other hand, U(III) was not observed in the 2-PrOH system because

(CH3)2COH radicals, formed by hole/HO attack to 2-PrOH and able to produce U(III)

25
(E0 from 1.8 to 1.39 V) [45] have been proposed not to be directly involved in the

U(VI) reduction mechanism, which is mediated only by eCB. This is a distinctive

feature of the 2-PrOH system, probably due to the low reaction rate of U(VI) with

organic radicals [18], as proposed also for the homogeneous system [10]. It is necessary

to remember that the TiO2 nanoparticles are synthesized in a large amount of 2-PrOH

(around 1 M); when HCOOH is absent (Fig. 4), U(III) was not detected, confirming

that, even at low pH, it cannot be formed by (CH3 )2COH. In contrast, Fig. 5 shows the

ability of CO2 to reduce U(VI) up to U(III) in the presence of HCOOH. No reports

exist on the values of the rate constants of the reaction of U(VI) or U(IV) with COO,

but it can be suggested that they are higher than the constants of the reaction of these

species with organic radicals (e.g., (CH3)2COH), as expected from: 1) the higher

reduction potential of COO compared with that of organic radicals [45], and 2) the

higher affinity between the positive U(VI) or U(IV) ions and COO. All this discussion

reinforces that, up to now, only HCOOH has proven to be able to promote reduction of

U(IV) to U(III).

According to the above reported rate constants for the reactions of HO in water with

HCOOH and U(IV), and the lifetime of U(V), at [HCOOH] 0.01 M, HCOOH would

be oxidized in preference to the uranium reduced species. At smaller [HCOOH],

U(V)/U(IV)/U(III) oxidation by HO would be preferred to HCOOH oxidation;

however, HCOOH could be preferentially oxidized by hVB+ due to a better adsorption

on TiO2 compared with the cationic uranium species. Besides, U(IV)/U(V) oxidation by

hVB+/HO yields U(V)/U(VI), which can then react with eCB-, with no net change.

26
A possible competition of the reaction of carboxyl radicals with U(IV) is injection of

eCB; as can be estimated from the values reported in [61], the second order rate constant

for the reaction of a carboxyl radical with TiO2 to give CO2 plus eCB is 4 104 M1 s1,

while the combination rate between two carboxyl radicals is around 4-8 108 M1 s1

[62]. Given the short lifetime of this radical, this reaction can be discarded. The results

of the present work indicate that COOH/COO would react in preference with U(IV)

giving U(III); on the contrary, the blue color of eCB should have been observed [63].

Injection of electrons into the TiO2 CB by U(III) (Eq. (13)) would be

thermodynamically possible. However, this reaction does not take place, as the

development of the U(III) signal in Fig. 5 was observed without the simultaneous

increase of the broad band of trapped electrons around 600 nm (with a characteristic

blue color) [11].

U(III) U(IV) + eCB (13)

A much faster reaction of hBV+ with U(III) than with HCOOH would imply that this

species could not be detected, as the remaining eCB cannot reduce U(IV) to U(III); the

net results would be similar to Eq. (13). As stated before, as the characteristic blue color

of trapped eCB was not observed, a significant oxidation of U(III) by hBV+ in the studied

conditions can be discarded.

Reduction of H+ to H2 by U(III) (Eq. (13)) [27] is also negligible at the working pH,

as U(III) half-life for this decay is at least 755 h [37].

27
U(III) + H+ U(IV) + H2 (14)

U(III), together with all uranium reduced forms could be spectrophotometrically seen

at acid pH also in the system in the presence of HCOOH and absence of TiO2 (Fig. 6),

but the amount of U(III) was very small. U(III) formation during photochemical

reactions has not been previously reported and represents a significant contribution to

the elucidation of both, homogeneous and heterogeneous photochemical mechanisms.

Comparing now the efficiency of the HP process with 2-PrOH [9] and with HCOOH,

despite the negative effect of HCOOH at high concentrations (Figs. 1 and 2), this

electron donor seems to be more beneficial; moreover, very small amounts of HCOOH

(1 M, i.e. HCOOH/U(VI) molar ratio = 4) cause a very high removal initial rate, while

2-PrOH needs much higher 2-PrOH/U(VI) molar ratios ( 4 103). The higher

efficiency of HCOOH can be attributed to the higher adsorption of HCOOH over TiO2,

which promotes a more efficient scavenging of hVB+ and a faster oxidation by HO (Eq.

(11)), and to the high COO reducing power. These results are in agreement with those

found by Amadelli et al. [1].

As expected, a less efficient photocatalytic U(VI) reduction took place in aerated

suspensions (Fig. 2) due to the competition between O2 and uranyl for eCB and/or the

reducing radicals, and because O2 can easily reoxidize U(IV)/U(V) to U(VI); this is

consistent with previous results [3,4,9]. However, it is important to remark that the

effect of O2 is somewhat counteracted because it is partially used in the oxidation of

formic acid. The difference between U(VI) removal under air bubbling (10%) or with

28
the reactor open to air (50%) can be ascribed only to the smaller oxygen concentration

present under the last condition. Of course, the efficiency of U(VI) reduction increases

considerably when O2 is totally replaced by direct N2 bubbling in the suspension.

The lower yield of the HP reaction for QAN compared with QP (cf. Figs. 1 and 2) is

explained by the formation of soluble U(IV)-acetate complexes. hVB+ or HO will

preferentially react with HCOOH instead of acetate due to a more suitable reduction

potential (cf. E0(CH2CO2H, H+/CH3CO2 H) = 1.8 V and E0(HCOO, H+/HCOOH) = 1.3

V) [64] and also because HCOOH is in excess with respect to acetate, even at the lowest

HCOOH concentration (0.00025 M), where acetate concentration is twice HCOOH.

Similarly to the 2-PrOH system either in the absence [10] or in the presence of TiO2

[9], the yield in the HCOOH system was higher with the Q photoreactor (Fig. 1), and

this can be explained by the more energetic radiation transmitted by the material, the

higher amount of photons arriving to the solution, and the higher U(VI) absorption. A

change in the mechanism due to the different wavelength range is not expected.

Effects of nitrate (QAN) were considered negligible. In the 2-PrOH system, as said,

no significant differences were observed using uranyl nitrate or perchlorate. NO3

reduction is not possible by eCB in the absence of HCOOH and, although it could be

reduced by COO, our evidences point out that U(VI) is preferentially reduced. Effects

of ClO4 will be also considered negligible. Although Cohen and Carnall [20] indicate

that ClO4 can be slowly reduced by U(III), Sat et al. [19] report no reaction and, in

any case, ClO4 can be considered inert toward U(III) in the timescale of the present

29
experiments. On the other hand, ClO4 was found inert in the HP uranyl/2-PrOH system

[9].

5. Conclusions

The UV heterogeneous photocatalysis of the uranyl ion in the presence of HCOOH

under anoxic conditions is a very efficient process to remove uranyl from water by

conversion to U(IV). However, the counterion of the uranyl salt has effect on the

efficiency of the reaction, and the best removal conditions have been established, the

perchlorate being the best system (and probably the nitrate, not tested here). However,

excess of HCOOH should be avoided because, despite a significant initial conversion,

an important reoxidation took place, which prevents U(VI) removal. This can be

explained by the formation of the soluble and easily reoxidizable U(IV)-formate

complex, and of soluble U(III). Less HCOOH concentrations than those of 2-PrOH are

required to obtain very good removal efficiencies. In addition, CO2 is the only product

of the reaction, in contrast with 2-PrOH, which is transformed into acetone and can be

environmentally harmful.

The experiment with nanoparticles at very low pH values (< 2) allowed to detect and

confirm spectrophotometrically the generation of U(III) in the presence of a high

HCOOH concentrations. It was proved that U(III) is formed at a large extent only in the

presence of TiO2 and HCOOH. This is the first time that the formation of U(III) in a

photochemical U(VI) system has been reported.

30
Acknowledgments

This work was performed as part of Agencia Nacional de Promocin Cientfica y

Tecnolgica PICT-512 and PICT-0463 projects. To A.G. Leyva and G. Custo for XRD

analysis and TXRF measurements, respectively.

References

[1] R. Amadelli, A. Maldotti, S. Sostero, V. Carassiti, Photodeposition of uranium

oxides onto TiO2 from aqueous uranyl solutions, J. Chem. Soc. Faraday Trans.

87 (1991) 32673273.

[2] C. Cerrillos, D.F. Ollis, Photocatalytic reduction and removal of uranium from a

uranium-EDTA solution, J. Adv. Oxid. Technol. 3 (1998) 167173.

[3] J. Chen, D.F. Ollis, W.H. Rulkens, H. Bruning, Photocatalyzed deposition and

concentration of soluble uranium (VI) from TiO2 suspensions, Col. Surf. A:

Physicochem. Eng. Aspects, 151 (1999) 339349.

[4] V. Eliet, G. Bidoglio, Kinetics of the Laser Induced Photoreduction of U(VI) in

aqueous suspensions of TiO2 particles, Environ. Sci. Technol., 32 (1998) 3155

3161.

[5] M. Bonato, G.C. Allen, T.B. Scott, Reduction of U(VI) to U(IV) on the surface

of TiO2 anatase nanotubes, Micro & Nano Lett., 3 (2008) 5761.

31
[6] M.I. Litter, Heterogeneous photocatalysis. Transition metal ions in

photocatalytic systems Appl. Catal. B: Environ., 23 (1999) 89114.

[7] M.I. Litter, Treatment of chromium, mercury, lead, uranium and arsenic in

water by heterogeneous photocatalysis Adv. Chem. Eng., 36 (2009) 3767.

[8] M.I. Litter, N. Quici, New advances of heterogeneous photocatalysis for

treatment of toxic metals and arsenic, in: B.I. Kharisov, O.V. Kharissova, H.V.

Rasika Dias (Eds.), Nanomaterials for Environmental Protection, John Wiley &

Sons, Hoboken, New Jersey, in press.

[9] V.N. Salomone, J.M. Meichtry, G. Zampieri, M.I. Litter, New insights in the

heterogeneous photocatalytic removal of U(VI) in aqueous solution in the

presence of 2-propanol Chem. Eng. J., in press.

[10] V.N. Salomone, J.M. Meichtry, G. Schinelli, A.G. Leyva, M.I. Litter,

Photochemical reduction of U(VI) in aqueous solution in the presence of 2-

propanol, J. Photochem. Photobiol. A., 277 (2014) 1925.

[11] C. Kormann, D.W. Bahnemann, M.R. Hoffmann, Preparation and

characterization of quantum-size titanium dioxide, J. Phys. Chem., 92 (1988)

51965201.

[12] T.M. Florence, Y. Farrar, Spectrophotometric determination of uranium with

4-(2-pyridylazo) resorcinol, Anal. Chem., 35 (1963) 16131616.

32
[13] T. Arai, Y. Wei, M. Kumagai, An efficient elution method of tetravalent

uranium from anion exchanger by using formic acid solution, J. Alloys Comp.,

451 (2008) 400402.

[14] H.D. Burrows, T.J. Kemp, The photochemistry of the uranyl ion, Chem. Soc.

Rev., 3 (1974) 139165.

[15] J.T. Bell, R.E. Biggers, The absorption spectrum of the uranyl ion in

Perchlorate. Part I. Mathematical Resolution of the Overlapping Band Structure

and Studies of the Environmental Effects, J. Molec. Spectrosc., 18 (1965) 247

275.

[16] J.T. Bell, R.E. Biggers, The absorption spectrum of the uranyl ion in

Perchlorate. Part II. The Effect of hydrolysis on the resolved spectral bands, J.

Molec. Spectrosc., 22 (1967) 262271.

[17] L.J. Heidt, K.A. Moon, Evidence for pentavalent uranium as an intermediate in

the reaction in water between photoactivated uranyl ions and sucrose and

closely related substances, and quantum yields for these reactions, J. Am.

Chem. Soc., 75 (1953) 58035809.

[18] Y. Katsumura, H. Abe, T. Yotsuyanagi, K. Ishigure, Photochemical reactions

of uranyl ion in nitric acid - quantum yields of photoemission and

photoreduction with ethanol, J. Photochem. Photobiol. A. 50 (1989) 183197.

33
[19] A. Sat, S. Suzuki, Studies of the behaviour of trivalent uranium in an aqueous

Solution. II. Absorption spectra and ion exchange behaviour in various acid

solutions, Bull. Chem. Soc. Japan., 41 (1968) 26502656.

[20] D. Cohen, W.T. Carnall, Absorption spectra of uranium(III) and uranium(IV)

in DClO4 solution, J. Phys. Chem., 64 (1960) 19331936.

[21] S.N. Guha, P.N. Moorthy, K.N. Rao, Radiation induced redox reactions in the

U(VI)-U(IV) system in aqueous H2SO4 solutions, Radiat. Phys. Chem., 29

(1987) 425428.

[22] D. Cohen, The preparation and spectrum of U(V) in aqueous solutions, J.

Inorg. Nucl. Chem., 32 (1970) 35253530.

[23] C. Miyake, T. Kondo, S. Imoto, H. Ohya-Nishiguchi, Direct evidence of

uranium(V) intermediates by electron spin resonance in photo- and electrolytic

reduction processes of uranyl complexes in organic solutions, J. Less.

Common. Met., 122 (1986) 313317.

[24] J.T. Bell, H.A. Friedman, M.R. Billings, Spectrophotometric Studies of

dioxouranium(V) in aqueous media I. The perchlorate medium, J. Inorg. Nucl.

Chem., 36 (1974) 25632567.

[25] Y. Di Iorio, E. San Romn, M.I. Litter, M.A. Grela, Photoinduced reactivity of

strongly coupled TiO2 ligands under visible irradiation: an examination of an

alizarin Red@TiO2 nanoparticulate system, J. Phys. Chem. C., 112 (2008)

1653216538.

34
[26] C.K. Jrgensen, Absorption spectra of red uranium(III) chloro complexes in

strong hydrochloric acid, Acta Chem. Scand., 10 (1956) 15031505.

[27] S.A. Kulyukhin, N.B. Mikheev, A.N. Kamenskaya, N.A. Konovalova, I.A.

Rumer, Physicochemical properties of uranium in lower oxidation states

Radiochemistry, 48 (2006) 535551.

[28] J. Drodyski, Tervalent uranium compounds, Coord. Chem. Rev., 249

(2005) 23512373.

[29] J. Drodyski, K. Schwochau, H.-J. Schenk, Spectroscopic, magnetic and X-

ray powder diffraction studies on U(III) and Np(III) formates, J. Inorg. Nucl.

Chem., 43 (1981) 18451848.

[30] J. Drodyski, J.G. Conway, Low-temperature absorption spectrum of

uranium (3+) formate energy level scheme of uranium (3+), J. Chem. Phys. 56

(1972) 883891.

[31] S.J. Formosinho, H.D. Burrows, M.G. Miguel, M.E.D.G. Azenha, I.M.

Saraiva, A.C.D. N. Ribeiro, I.V. Khudyakov, R.G. Gasanov, M. Bolte, M.

Sarakha, Deactivation processes of the lowest excited state of [UO2(H2O)5]2+ in

aqueous solution, Photochem. Photobiol. Sci., 2 (2003) 569575.

[32] J. Selbin, J.D. Ortego, The chemistry of U(V), Chem. Rev., 69 (1969) 657

671.

35
[33] D.M.H. Kern, E.F. Orlemann, The Potential of the Uranium (V), Uranium (VI)

Couple and the Kinetics of Uranium (V) Disproportionation in Perchlorate

Media, J. Am. Chem. Soc., 71 (1949) 21022106.

[34] A. Ekstrom, Kinetics and Mechanism of the Disproportionation of

Uranium(V), Inorg. Chem., 13 (1974) 22372241.

[35] E. Atinault, V. De Waele, J. Belloni, C. Le Naour, M. Fattahi, M. Mostafavi,

Radiolytic Yield of UIV Oxidation into UVI: A New Mechanism for UV

Reactivity in Acidic Solution, J. Phys. Chem. A., 114 (2010) 20802085.

[36] A.J. Elliot, S. Padamshi, J. Pika, Free-radical redox reactions of uranium ions

in sulphuric acid solutions, Can. J. Chem., 64 (1986) 314320.

[37] Y. Kobayashi, A. Saito, Autoxidation and metal surface catalysed oxidation of

uranium(III) in acid solutions, J. Inorg. Nucl. Chem., 41 (1979) 15951599.

[38] S.T. Martin, H. Herrmann, M.R. Hoffmann, Time-resolved microwave

conductivity. Part 2.Quantum-sized TiO2 and the effect of adsorbates and

light intensity on charge-carrier dynamics, J. Chem. Soc., Faraday Trans., 90

(1994) 33233330.

[39] L. Martinot, J. Fuger, The actinides in Standard potentials in aqueous solution,

in: A.J. Bard, R. Parsons, J. Jordan (Eds.), Marcel Dekker, New York, 1985,

pp. 631674.

36
[40] M. Duflo-Plissonnier, K. Samhoun, Etude de la reduction U/III/-U/O/ en

solution aqueuse par polarographie radiochimique, Radiochem. Radioanal.

Lett., 12 (1972) 131138.

[41] C.J. Evans, G.P. Nicholson, D.A. Faith, M.J. Kan, Photochemical removal of

uranium from a phosphate waste solution, Green Chem., 6 (2004) 196197.

[42] M.R. Hoffmann, S.T. Martin, W. Choi, D.W. Bahnemann, Environmental

applications of semiconductor photocatalysis, Chem. Rev. 95 (1995) 6996.

[43] G.V. Buxton, C.L. Greenstock, W.P. Helman, A.B. Ross, Critical review of

rate constants for reactions of hydrated electrons, hydrogen atoms and hydroxyl

radicals (OH/O-) in aqueous solution, J. Phys. Chem. Ref. Data 17 (1988), 513-

886.

[44] S. Gordon, J.C. Sullivan, A.B. Ross, Rate constants for reactions of radiation-

produced transients in aqueous solutions of actinides, J. Phys. Chem. Ref. Data, 15

(1986) 13571367.

[45] P.J. Wardman, Reduction potentials of one-electron couples involving free

radicals in aqueous solution, J. Phys. Chem. Ref. Data., 18 (1989) 16371755.

[46] E. Rabinowich, Photochemistry of uranyl compounds, Pergamon Press,

Oxford, U.K., 1954, pp. 296.

[47] G.E. Heckler, A.E. Taylor, C. Jensen, D. Percival, R. Jensen, P. Fung, Uranyl

sensitized photodecomposition of organic acids in solution, J. Phys. Chem., 67

(1963) 16.

37
[48] D. Greatorex, R.J. Hill, T.J. Kemp, T.J. Stone, Electron spin resonance studies

of photo-oxidation by metal ions in rigid media at low temperatures. Part 4.

Survey of photo-oxidation by the uranyl ion J. Chem. Soc., Faraday Trans. 1,

68 (1972) 20592076.

[49] M. Goldstein, J.J. Barker, T. Gangwer, A photochemical technique for

reduction of uranium and subsequently plutonium in the Purex process.

September 1976, informal report, Department of Applied Science, Brookhaven

National laboratory, Associated Universities, Inc.

[50] A.G. Brits, R. Van Eldik, J.A. Van Der Berg, The photolysis of the uranyl

formic acid/formate system in acidic aqueous solution, Inorg. Chim. Acta., 30

(1978) 1722.

[51] J. Hu, X. Zhang, D. Yunfu, Z. Zhihong, H. Xu, Studies of the photochemical

reduction of uranyl nitrate in aqueous solution, J. Less-Common Metals, 122

(1986) 287 294.

[52] D.M. Roundhill, Photochemistry and Photophysics of Lanthanide and Actinide

Complexes, in Photochemistry and Photophysics of Metal Complexes. Springer

Science + Business Media, New York, 1994, pp. 303320.

[53] T.M. McCleskey, T.M. Foreman, E.E. Hallman, C.J. Burns, N.N. Sauer,

Approaching Zero Discharge in Uranium Reprocessing: Photochemical

Reduction of Uranyl Environ. Sci. Technol., 35 (2001) 547551.

38
[54] C. Lucks, A. Rossberg, S. Tsushima, H. Foerstendorf, K. Fahmy, G. Bernhard,

Formic acid interaction with the uranyl(VI) ion: structural and photochemical

characterization, J. Chem. Soc. Dalton Trans., 42 (2013) 1358413589.

[55] G.V. Buxton, R.M. Sellers, Acid dissociation constant of the carboxyl radical.

Pulse radiolysis studies of aqueous solutions of formic acid and sodium

formate, J. Chem. Soc., Faraday Trans. 1, 69 (1973) 555559.

[56] I. Grenthe, J. Drodyski, T. Fujino, E.C. Buck, T.E. Albrecht-Schmitt and

S.F. Wolf, Uranium, in: L.R. Morss, N.M. Edelstein, J. Fuger, J.J. Katz, The

Chemistry of the Actinide and Transactinide Elements, fourth ed., Springer,

Dordrecht, 2011.

[57] N.W. Alcock, Uranyl oxalate complexes. Part I. Preparation and crystal and

molecular structure of ammonium uranyl trioxalate, J. Chem. Soc., Dalton

Trans., (1973) 16141620.

[58] D. Golub, H. Cohen, D. Meyerstein, Kinetics and mechanism of single electron

oxidations of the tervalent uranium ion, U3+ (aq), by free radicals in aqueous

solutions, J. Chem. Soc., Dalton Trans., (1985) 641644.

[59] G.R. Dey, K.N.R. Nair, K.K. Pushpa, Photolysis studies on HCOOH and

HCOO in presence of TiO2 photocatalyst as suspension in aqueous medium, J.

Nat. Gas Chem., 18 (2009) 5054.

39
[60] N.L. Banik, B. Brendebach, C.M. Marquardt, Investigations of actinides in the

context of final disposal of high-level radioactive waste: trivalent actinides in

aqueous solution, J. Radioanal. Nucl. Chem., 300 (2014) 177183.

[61] L.L. Perissinotti, M.A. Brusa, M.A. Grela, Yield of carboxyl anion radicals in

the photocatalytic degradation of formate over TiO2 particles, Langmuir, 17

(2001) 84228427.

[62] P. Neta, R.E. Huie, A. Roos, Rate constants of inorganic radicals in aqueous

solution, J. Phys. Chem. Ref. Data, 17 (1988) 1027-1284.

[63] D. Bahnemann, A. Henglein, J. Lilie, L. Spanhel, Flash photolysis observation

of the absorption spectra of trapped positive holes and electrons in colloidal

TiO2, J. Phys. Chem., 88 (1984) 709711.

[64] D. Yu, A. Rauk, D.A. Armstrong, Radicals and ions of formic and acetic acids:

an ab initio study of the structures and gas and solution phase thermochemistry,

J. Chem. Soc. Perkin Trans., 2 (1994) 22072215.

40
Fig. 1. Temporal evolution of normalized U(VI) concentration in the presence of

HCOOH during HP experiments with uranyl perchlorate. Conditions: setup I, QP and

GP, [U(VI)]0 = 0.25 mM, [TiO2] = 1 g L1, N2 (0.5 L min1), pH 3, T = 25 C. Q

photoreactor: > 230 nm, max = 365 nm, q 0n,p/V = 121 einstein s1 L1. G

photoreactor: > 310 nm, max = 365 nm, q0n,p/V = 44 einstein s1 L1. Dotted lines are

fittings of experimental points to Eq.(1), with the exception of GP with 1 M HCOOH,

which was fitted up to 5 min, and experiments without TiO2, where dotted lines are only

for a better visualization of the experimental points.

Fig. 2. Temporal evolution of normalized U(VI) concentration in the presence of

HCOOH during experiments with TiO2 with uranyl acetate (QAN conditions).

Experiments were run up to 120 min but no changes were observed after 30 min.

Conditions of Fig. 1. Dotted lines are fittings of experimental points to Eq. (1) with the

exception of the curves with 0.1 and 1 M HCOOH, which were fitted up to 5 min.

Fig. 3. Spectra of the initial and final filtered solutions of a HP experiment for QAN in

the presence of 1 M HCOOH. Conditions of Fig. 2. Spectra of the initial and final

filtered solutions of a HP experiment for QAN in the presence of 1 M HCOOH.

Conditions of Fig. 1. Insets: detailed views of the 200-350 nm and 350-500 nm regions.

Fig. 4. Temporal evolution of the spectrum of a uranyl perchlorate solution in the

presence of TiO2 nanoparticles and absence of HCOOH. Conditions: setup II, [U(VI)]0

= 1 mM, [TiO2] = 0.010 M, pH 1.80, N2 atmosphere, E0 = 3700 W cm2, full emission

spectrum. The spectrum in the absence of U(VI) is also included. Inset: detailed view of

the 350-500 nm region.


Fig. 5. Temporal evolution of the spectrum of a uranyl perchlorate solution in the

presence of TiO2 nanoparticles and HCOOH. Conditions: setup II, [U(VI)]0 = 1 mM,

[TiO2] = 0.010 M, [HCOOH] = 1 M, pH 1.3, N2 atmosphere, E0 = 3700 W cm2, full

emission spectrum. Inset: detailed view of the 350-600 nm region.

Fig. 6. Temporal evolution of the spectrum of a uranyl perchlorate solution in the

presence of HCOOH (no TiO2 nanoparticles) up to 90 min of UV-Vis irradiation and

after 90 min of UV-Vis irradiation plus 30 min of exposure in the dark to the air.

Conditions: setup II, [U(VI)]0 = 1 mM, [HCOOH] = 1 M, pH 1.9, N2 atmosphere, E0 =

3700 W cm2, full emission spectrum. Inset: detailed view of the 350-600 nm region.

Scheme 1. Latimer diagram for different uranium species vs. SHE [27,38-40].

You might also like